Enhanced Performance of Perovskite CH3NH3PbI3 Solar Cell by

May 26, 2015 - Sequential deposition is a widely adopted method to prepare CH3NH3PbI3 on mesostructured TiO2 electrode for organic lead halide perovsk...
0 downloads 11 Views 2MB Size
Subscriber access provided by NEW YORK UNIV

Article

Enhanced Performance of Perovskite CH3NH3PbI3 Solar Cell by Using CH3NH3I as Additive in Sequential Deposition Yian Xie, Feng Shao, Yaoming Wang, Tao Xu, Deliang Wang, and Fuqiang Huang ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.5b02705 • Publication Date (Web): 26 May 2015 Downloaded from http://pubs.acs.org on May 30, 2015

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Applied Materials & Interfaces is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Enhanced Performance of Perovskite CH3NH3PbI3 Solar Cell by Using CH3NH3I as Additive in Sequential Deposition Yian Xie,a,b Feng Shao,a,c Yaoming Wang,a Tao Xu,d,* Deliang Wang,c and Fuqiang Huanga,b,* a

CAS Key Laboratory of Materials for Energy Conversion, Shanghai Institute of Ceramics,

Chinese Academy of Sciences, Shanghai 200050, P. R. China b

State Key Laboratory of Rare Earth Materials Chemistry and Applications, College of

Chemistry and Molecular Engineering, Peking University, Beijing 100871, P. R. China c

Hefei National Laboratory for Physical Sciences at the Microscale, University of Science and

Technology of China, Hefei 230026, P. R. China d

Department of Chemistry and Biochemistry, Northern Illinois University, DeKalb, Illinois

60115, USA. Keywords: Perovskite, solar cell, sequential solution deposition, recombination, CH3NH3I.

Abstract : Sequential deposition is a widely adopted method to prepare CH3NH3PbI3 on mesostructured TiO2 electrode for organic lead halide perovskite solar cells. However, this method often suffers from the uncontrollable crystal size, surface morphology and residual PbI2

ACS Paragon Plus Environment

1

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 20

in the resulting CH3NH3PbI3, which are all detrimental to the device performance. We herein present an optimized sequential solution deposition method, by introducing different amount of CH3NH3I in PbI2 precursor solution in the first step, to prepare CH3NH3PbI3 absorber on mesoporous TiO2 substrates. The addition of CH3NH3I in PbI2 precursor solution can affect the crystallization and composition of PbI2 raw films, resulting in the variation of UV-Vis absorption and surface morphology. Proper addition of CH3NH3I not only enhances the absorption, but also improves the efficiency of CH3NH3PbI3 solar cells from 11.13% to 13.37%. Photoluminescence spectra suggest that the improvement of device performance is attributed to the decrease of recombination rate of carriers in CH3NH3PbI3 absorber. This current method provides a highly repeatable route for enhancing the efficiency of CH3NH3PbI3 solar cell in sequential solution deposition method.

Introduction In the last five years, organic−inorganic hybrid perovskites (e.g. CH3NH3PbI3) have gained extensive attention around the world and high power conversion efficiency close to 20%1-2 is achieved by these perovskites-based solar cells, which is comparable to commercialized c-Si solar cells and state of art Cu(In,Ga)(S,Se)2 thin film photovoltaic solar cells.3 CH3NH3PbI3 perovskite affords several important advantages such as high absorption coefficient (> 104 cm1 4-5

),

highly mobile electron and holes,6 long carrier transport diffusion length (100 to 1000

nm),5,

7-8

suitable and tunable direct band gap by controlling the chemical composition,9-11

making it an outstanding light harvester and an excellent hole transport material. Furthermore, CH3NH3PbI3 solar cells can be fabricated via simple and low cost chemical solution methods, which has great potential in large-scale production.

ACS Paragon Plus Environment

2

Page 3 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Presently, the most adapted and traditional chemical solution deposition methods can be classified into two categories, namely, one step deposition and sequential deposition methods. As for one step deposition method, Cl is powerful additive in precursor solution with the forms including PbCl212-13 or CH3NH3Cl.14 The addition of Cl can slow down the crystallization rate during annealing process and thus control the morphology of CH3NH3PbI3-xClx thin films. Although high efficiency CH3NH3PbI3-xClx solar cells have been successful constructed by one step deposition method, the atomic ratio of Cl in the perovskite as well as the reaction mechanism that leads to its formation and crystallization are still under debate.15-17 In a typical sequential deposition process, PbI2 layer or film is firstly deposited on compact TiO2 layer or mesoporous TiO2 layer from its N,N-dimethylformamide (DMF) solution, followed by immersing the PbI2 film in a CH3NH3I (MAI) 2-propanol solution to form the desired CH3NH3PbI3. However, two main problems are found in sequential deposition. Firstly, the crystal size and surface morphology of CH3NH3PbI3 are uncontrollable, which is detrimental to the device reproducibility. Secondly, the residual PbI2 in CH3NH3PbI3 thin films deteriorates the reproducibility of device performance as well. Although some work tackled this problem, including depositing a high coverage CH3NH3PbI3 capping layer on mesoporous TiO2 films by spin-coating the PbI2 layer twice,18 or using dimethyl sulfoxide (DMSO) as the solvent for better film quality,19 a more reliable access to acquire CH3NH3PbI3 with less structural defects is still exceptionally desired. In this work, we developed an alternative approach to tackle this problem. We introduced a certain amount of MAI in the precursory PbI2 solution during the first step. The crystallization and surface morphology of CH3NH3PbI3 became very sensitive to the addition of MAI, while suitable amount of MAI can improve the absorption without disturbing the high coverage of the

ACS Paragon Plus Environment

3

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 20

capping layer. In comparison to the devices without the addition of MAI in the first step, power conversion efficiency is enhanced from 11% to 13% under optimum addition amount of the initial MAI. The improvement in device performance can be owing to the increase of shortcircuit current (Jsc) and slight increase of open-circuit voltage (Voc). The recombination rate of carriers is investigated by photoluminescence (PL) spectra. Experimental details Transparent conducting substrate and mesoporous TiO2 thin film. Fluorine-doped SnO2-coated transparent conducting glass substrate (FTO) was washed by ultrasonication with detergent firstly, then with deionized water, acetone and ethanol, and finally dried with dry nitrogen air. A 50-nm-thick TiO2 compact layer was then deposited on the substrates by sputtering at 0.6 Pa with 120 W for 30min. The mesoporous TiO2 film was prepared by spin-coating a 20 nm-sized TiO2-nanoparticulate paste (diluted in ethanol with a ratio of 1:5 by weight, OPV-Tech, DHSTPP3) for 4 times at 2500 rpm for 30 s, dried at 125 °C for 10 min, then heated at 500 °C for 30 min. Device fabrication. CH3NH3I (MAI) was synthesized by reacting 19.5 mL methylamine (40 wt% aqueous solution) and 32.3 mL hydroiodic acid (55-58 wt% aqueous solution) in an ice bath for 2 h with stirring. The product was obtained by rotary evaporation at 45 °C and cleaning with diethyl ether followed by drying in vacuum oven at 60 °C overnight.20 Five precursory solution were prepared by dissolution of 1 mmol PbI2 and 0, 0.1, 0.2, 0.3 or 0.4 mmol MAI in 1 mL DMF at 60 °C, which is labeled as PbI2-xMAI (x = 0, 0.1, 0.2, 0.3, and 0.4, respectively). The precursors were spin-coated on the mesoporous TiO2 films at 6000 rpm for 10 seconds, then heating on a hotplate at 100 °C for 10 minutes. Subsequently, the films were infiltrated in 2propanol and blow-dried. Perovskite films were formed by immersing these PbI2-xMAI films in

ACS Paragon Plus Environment

4

Page 5 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

a 10 mg/mL MAI 2-propanol solution for 25 minutes, the obtained perovskite films are labeled as MAPI-xMAI, accordingly. After being rinsed with 2-propanol, the perovskite films were annealed at 100 °C for 10 minute. The hole transport material (HTM) solution was prepared as followed: 52.8 mg 2,29,7,79-tetrakis-(N,N-di-p-methoxyphenyl-amine) -9,99-spirobifluorene (spiro-MeOTAD) was dissolved in 640 µL chlorobenzene and mixed with 10 µL of 500 mg/mL bis(trifluoromethane)sulfonimide lithium salt (LiTFSI) solution in acetonitrile and 14.4 µL 4tert-butylpyridine (TBP). The HTM solution was spin-coated on the perovskite-covered TiO2 electrodes at 2500 rpm for 30 seconds. For the counter electrode, a 100 nm-thick Au was deposited on the top of the HTM layer by a thermal evaporation and the active area was 0.07 cm2. Characterization. The crystallization and phase identification of the thin films were performed by X-ray diffraction (XRD Bruker D8 Focus) with a monochromatized source of Cu Kα1 radiation (λ = 0.15405 nm) at 1.6 kW (40 kV, 40 mA). UV-Vis absorbance spectrums were recorded on a Hitachi U-3010 spectrophotometer with a scanning velocity of 300 nm min-1. Raman spectra were collected on a Thermal Dispersive Spectrometer using a laser with an excitation wavelength of 532 nm at laser power of 10 mW. Top-view field emission scanning electron microscopy (FESEM) images were taken on ZEISS SUPRA 55 microscope. PL spectra were measured at 300 K on a fluorescence spectrophotometer (F-4600, Hitachi, Japan) with an excitation wavelength of 466 nm. Photocurrent density–voltage characteristics were measured using a Keithley Model 2440 source meter under AM 1.5 illumination. A 1000 W Oriel solar simulator was used as a light source and the power of the light was calibrated to one sun light intensity by using a NREL-calibrated Si cell (Oriel 91150). The scan rate was 57.5 mV s-1 and

ACS Paragon Plus Environment

5

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 20

the scan direction was from high to low voltage. The external quantum efficiency (EQE) was measured by a Newport QE system equipped with a 300 mW xenon light and a lock-in amplifier. Results and discussion XRD analysis was conducted to verify the crystallization of the PbI2-xMAI and final CH3NH3PbI3 layer on mesoporous TiO2 films. The XRD patterns of PbI2-xMAI films were shown in Figure 1a and the detailed information between 11.5° and 15° was displayed in Figure 1b. As for PbI2-0.1MAI and PbI2-0.2MAI, the main peak of CH3NH3PbI3 at 14.2° cannot be detected though a small amount of MAI was added in PbI2 precursor solution (Figure 1b), which may be resulted from the low crystallization of few CH3NH3PbI3 and the decomposition of CH3NH3PbI3 at high temperature (100 °C). On the other hand, the addition of small amount of MAI even led to better crystallization of PbI2, indicated by the increasing peak intensity of PbI2 at 12.7° in the range from PbI2-0MAI to PbI2-0.2MAI. The characteristic peak of CH3NH3PbI3 at 14.2° was observed in PbI2-0.3MAI, and this peak intensity increased in PbI2-0.4MAI. The peak intensity of PbI2 in these two samples decreased significantly compared with PbI2-0.2MAI, which should be owing to the conversion of PbI2 to CH3NH3PbI3. Besides, the peak intensity of CH3NH3PbI3 was much lower than that of PbI2 even in PbI2-0.4MAI, implying the low content of CH3NH3PbI3 in the films. The color change can be observed (Figure 1c) and the absorption was measured by UV-Vis spectra (Figure 1d). Not only the PbI2's characteristic absorbance between 400 nm and 520 nm increased, but also the absorbance between 520 nm and 780 nm from CH3NH3PbI3 enhanced gradually with increasing MAI. The enhancement of absorbance between 520 nm and 780 nm should be caused by the insertion of MAI in PbI2 and the formation of CH3NH3PbI3. Although there was no signal of CH3NH3PbI3 detected in XRD patterns of PbI20.1MAI and PbI2-0.2MAI, MAI should be retained in thin films.

ACS Paragon Plus Environment

6

Page 7 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure. 1 (a) XRD patterns and (b) detail XRD information from 11° to 15° 2θ value of PbI2xMAI on mesoporous TiO2 films with varying x value (x = 0-0.4). The main XRD peaks of FTO, PbI2 and CH3NH3PbI3 were labeled. (c) Images and (d) UV-Vis absorbance spectra of PbI2-xMAI thin films as a function of x value. The corresponding CH3NH3PbI3 films, denoted as MAPI-xMAI films, were formed by immersing these PbI2-xMAI samples in a solution of 10mg/mL MAI in 2-propanol, and their XRD patterns are collected in Figure 2a. There is a weak signal of PbI2 observed in the XRD of MAPI-0MAI, indicating the incomplete conversion of PbI2 to CH3NH3PbI3, a common problem in sequential deposition.19, 21 However, residual PbI2 was also found in other samples, suggesting that the conversion of PbI2 to CH3NH3PbI3 was not readily achieved, while low content of CH3NH3PbI3 was formed in PbI2-xMAI (x = 0.3 or 0.4) beforehand. Derived from this, the low reaction driving force could be the main reason leading to the residue of PbI2. Long dipping time was necessary to completely convert PbI2 to CH3NH3PbI3.18 UV-Vis spectra (Figure 2b) showed

ACS Paragon Plus Environment

7

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 20

that the absorbance possessed a rinsing tendency in visible light region (between 400 nm and 780 nm) from MAPI-0MAI to MAPI-0.3MAI. For MAPI-0.4MAI, it’s noteworthy that the absorbance in near-infrared region (> 800 nm) was much higher than the previous four samples, which might resulted from surface scattering and lattice defects in CH3NH3PbI3. Moreover, the absorbance between 400 nm and 500 nm was lower than that of MAPI-0.2MAI and MAPI0.3MAI, manifesting the deteriorative absorption property occurred in MAPI-0.4MAI.

Figure 2 (a) XRD patterns and (b) UV-Vis absorbance spectra of MAPI-xMAI thin films as a function of x value (x = 0-0.4). The further information of PbI2-xMAI films was investigated by Raman spectroscopy and the Raman spectra were shown in Figure S1a (see supporting information). For PbI2-0MAI, the peak at 149 cm-1 was indexed to TiO2 and the Raman peaks of PbI2 were located at 76, 100, 114 and 172 cm-1. With increasing CH3NH3I content in PbI2 precursor, the peak intensity of PbI2 at 172 cm-1 decreased, and one small peak around 86-88 cm-1 appeared from PbI2-0.2MAI to PbI20.4MAI. Figure S1b showed the Raman spectra of MAPI-0.2MAI and mesoporous TiO2 film. Comparing these two Raman spectra, the Raman signal from 70 to 130 cm-1 should be assigned to the CH3NH3PbI3. One small peak at 87 cm-1 was observed in MAPI-0.2MAI, and the position of this peak was consistent with the peak in Figure S1a. So the peak around 86-88 cm-1 in Figure

ACS Paragon Plus Environment

8

Page 9 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

S1a should be the signal of CH3NH3PbI3. As for PbI2-0.2MAI, although on signal of CH3NH3PbI3 was found in XRD, the signal of CH3NH3PbI3 was detected in Raman spectra. The surface morphology and roughness of the PbI2-xMAI films were measured by FESEM and AFM. The top view FESEM images were shown in Figure S2. For PbI2-0MAI, the grain size was very small and almost no change on grain size was observed when x ≤ 0.2. When x ≥ 0.3, small dendritic grains was formed on the top of mesoporous TiO2, and the grain size increased with increasing x value, even some mesoporous TiO2 was exposed in PbI2-0.4MAI. As shown in Figure S3, the root mean square (RMS) of surface roughness were 14.02, 8.15, 8.67, 59.31 and 74.29 nm for PbI2-xMAI (x = 0, 0.1, 0.2, 0.3 and 0.4) films respectively. PbI2-0.1MAI and PbI20.2MAI films were relatively smoother than regular PbI2 film, while the surface roughness of PbI2-0.3MAI and PbI2-0.4MAI increased drastically. The increasing roughness of PbI2-0.3MAI and PbI2-0.4MAI were caused by the formation and crystal growth of CH3NH3PbI3 phase on surface. The surface morphology and roughness of the initial PbI2-xMAI exhibited a significant influence on the surface morphology of the subsequent MAPI-xMAI. The top view morphologies of CH3NH3PbI3 films prepared from different PbI2-xMAI precursors were characterized by FESEM and their images were shown in Figure 3a-e. The compact and holefree capping layers were achieved on MAPI-0MAI, MAPI-0.1MAI and MAPI-0.2MAI films through sequential deposition. The morphologies of CH3NH3PbI3 films based on MAPI-0MAI, MAPI-0.1MAI and MAPI-0.2MAI films were quite similar to each other and their crystal sizes were around 100 nm. In contrast, the MAPI-0.3MAI and MAPI-0.4MAI samples fabricated from these high RMS PbI2-xMAI films had lots of pinholes and voids, and the grain size of the films increased to 150-200 nm and 200-300 nm respectively. Thus, to achieve high surface coverage, the amount of MAI used in the first step should be kept in the relatively low region. The cross-

ACS Paragon Plus Environment

9

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 20

sectional morphology of MAPI-0.2MAI was shown in Figure 3f. The thickness of TiO2 scaffold layer and CH3NH3PbI3 capping layer were about 220 nm and 100 nm respectively.

Figure 3 Top view FESEM images of (a) MAPI-0MAI, (b) MAPI-0.1MAI, (c) MAPI-0.2MAI, (d) MAPI-0.3MAI, (e) MAPI-0.4MAI and cross sectional FESEM image of (f) MAPI-0.2MAI. The PL spectra were effective in exploring the recombination properties of light-excited electrons and holes in semiconductors. Figure 4 showed the PL spectra of varied MAPI-xMAI samples. All samples' emission peaks located at 764 nm (excitation wavelength: 446 nm) and were consistent with previous reports of emission from CH3NH3PbI318, 22-23 and the peak position of the emission was consistent among all of the samples. However, their PL intensities varied a lot and exhibited a decrease tendency from MAPI-0MAI to MAPI-0.2MAI and an increase tendency from MAPI-0.2MAI to MAPI-0.4MAI. The MAPI-0.2MAI showed the lowest peak intensity, a strong indication for low recombination and thus prospectively better photovoltaic

ACS Paragon Plus Environment

10

Page 11 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

performance.24 In contrast, MAPI-0.4MAI exhibited the highest PL signal and thus a higher recombination rate of carriers in comparison to other samples.

Figure 4 Steady-state PL spectra for MAPI-xMAI thin films with varying x value (x = 0-0.4). In current−voltage (J-V) curve measurement, the curve shape can be affected by scan rate significantly, as shown in Figure S4 and Table S1. In order to obtain conventional J-V curve and avoid long time radiation, scan rate of 57.5 mV s-1 was applied in J-V curve measurement. Figure 5a showed the typical J-V curves of the MAPI-xMAI-based devices and the device performance parameters were summarized in Table 1. The MAPI-0MAI-based cell exhibited a short-circuit photocurrent density (Jsc) of 16.15 mA cm-2, open-circuit voltage (Voc) of 1.015 V, and fill factor (FF) of 67.9%, yielding a power conversion efficiency of 11.13%. The efficiency increased to 11.92% and 13.37% for the MAPI-0.1MAI and MAPI-0.2MAI cells, respectively. Considering the similar FF values, the efficiency enhancement was mainly contributed by the higher Jsc and Voc. The Jsc for MAPI-0.1MAI and MAPI-0.2MAI increased to 16.90 and 18.58 mA cm-2 respectively, which was due to their stronger light absorption than MAPI-0MAI cell as shown in Figure 2b. The slight increase of Voc from 1.030 V in the MAPI-0.1MAI cell to 1.046 V in the

ACS Paragon Plus Environment

11

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 20

MAPI-0.2MAI cell should be owing to the less radiative recombination in the absorber of the MAPI-0MAI cell as manifested in Figure 4.1, 25 On the other hand, excessive addition of MAI (x = 0.3 and 0.4) in the first step gave adverse impact on device performance, as evidence by the sharp drop of Jsc and FF for MAPI-0.3MAI and MAPI-0.4MAI due to the high RRMS and low coverage of CH3NH3PbI3 capping layer, leading to more defective trap states for charge transport. According to the surface morphology, the lower device performance of MAPI-0.3MAI and MAPI-0.4MAI was due to the shunt leak through the pinholes in the CH3NH3PbI3, through which the subsequent hole-transport materials can be in direct contact with the TiO2. The averaged photovoltaic parameters extracted from J-V curves were collected in Figure S5, Voc, Jsc, FF and efficiency exhibited the similar tendency on x value in the MAPI-xMAI. The low error bars indicated the good repeatability of our device fabrication procedures, in favor of the high confidence level of this study. In addition, hysteresis is observed in MAPI-0.2MAI-based solar cell, as shown in Figure S6 and Table S2.

Figure 5 (a) Current–voltage (J-V) curve of the MAPI-xMAI-based (x = 0-0.4) solar cell under a standard AM 1.5 solar illumination at an intensity of 100 mW cm-2. (b) IPCE spectra of the MAPI-xMAI-based (x = 0-0.2) solar cell without any applied bias.

ACS Paragon Plus Environment

12

Page 13 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Table 1 Summary of MAPI-xMAI-based solar cell performance parameters with varying x values.

Sample

Voc (V)

Jsc (mA cm-2)

FF (%)

η (%)

MAPI-0MAI

1.015

16.15

67.9

11.13

MAPI-0.1MAI

1.030

16.90

68.5

11.92

MAPI-0.2MAI

1.046

18.58

68.8

13.37

MAPI-0.3MAI

0.943

9.11

50.8

4.38

MAPI-0.4MAI

0.929

7.67

53.7

3.83

The incident-photon-to-current conversion efficiency (IPCE) spectra (Figure 5b) revealed that the efficiency of MAPI-0MAI was higher than 80% in the range of 370-440 nm and dropped to lower than 60% in the range of 580-720 nm. The onset of photocurrent at 800 nm was consistent with the optical band gap of CH3NH3PbI3.26 The lower IPCE efficiency at long wavelength arose from inefficient charge extraction and/or light harvesting,27 resulting in the relative lower Jsc of solar cells. The MAPI-xMAIs with x = 0.1 and 0.2 exhibited a noticeable improvement of the efficiency over the whole region, especially in the long wavelength region between 500 and 720 nm for MAPI-0.2MAI. The peak IPCE value of MAPI-0.2MAI exceeded 85% around 410 nm and efficiency maintains above 60% in the range of 340-710 nm. Owing to the enhanced IPCE, the Jsc of MAPI-0.2MAI cells exhibited a pronounced improvement in photovoltaic performance. The Jsc for MAPI-0MAI, MAPI-0.1MAI and MAPI-0.2MAI, integrated from their IPCE spectra and the AM 1.5G solar photon flux,28 were 14.72, 15.37 and 17.08 mA cm-2 respectively. The slightly lower Jsc obtained from IPCE spectra, compared to the values measured from the J-V curve, can probably be attributed to the surface traps of the TiO2 layer.26, 29

ACS Paragon Plus Environment

13

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 20

Conclusion In summary, a new sequential deposition of CH3NH3PbI3 absorber on mesoporous TiO2 films was developed. Depending on the amount of MAI added in PbI2 precursor solution, the crystallization of PbI2 enhanced with low amount (x = 0.1 and 0.2) and CH3NH3PbI3 phase was formed with high amount (x = 0.3 and 0.4) in as-prepared PbI2-xMAI film. The PbI2 phase was expected to be removed completely with a certain amount addition of MAI, but PbI2 was still detected even in MAPI-0.4MAI. With low amount addition of MAI, the light absorption of CH3NH3PbI3 increased without obvious change of the surface morphology like grain size and flatness. What’s more, the recombination rate of carriers can be reduced, implying the better quality of CH3NH3PbI3. However, excessive addition of MAI not only increased roughness and lowered the surface coverage of capping layer, but also embodied higher recombination rate of carriers. Comparing with PbI2-0MAI, the efficiency improved from 11.13% to 13.37% with optimal addition amount of PbI2-0.2MAI. The improved device performance was attributed by enhanced Jsc and Voc, which should be the result of higher light absorption and better quality of CH3NH3PbI3 absorber. This method provided a new route for further improvement of device performance of CH3NH3PbI3 solar cell fabricated by sequential deposition.

Supporting Information: Raman spectra, top view FESEM images and AFM images of PbI2xMAI

film, average device parameters of MAPI-xMAI solar cell, scan rate-depended and scan

direction depended J-V curves are available in supporting information. This material is available free of charge via the Internet at http://pubs.acs.org. * Corresponding authors’ emails: [email protected] (T. Xu), [email protected]. (F. Huang)

ACS Paragon Plus Environment

14

Page 15 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

ACKNOWLEDGMENT The authors thank Prof. Yixin Zhao at Shanghai Jiao Tong University for kind suggestions and careful revision. This work was financially supported by National Science Foundation of China (Grants 91122034, 51125006, 61376056 and 61204072), and Science and Technology Commission of Shanghai (Grants 13JC1405700 and 14520722000). Tao Xu acknowledges the support from the U.S. National Science Foundation (CBET-1150617). REFERENCES 1.

Zhou, H.; Chen, Q.; Li, G.; Luo, S.; Song, T. B.; Duan, H. S.; Hong, Z.; You, J.; Liu, Y.;

Yang, Y., Photovoltaics. Interface Engineering of Highly Efficient Perovskite Solar Cells. Science 2014, 345, 542-546. 2.

Jeon, N. J.; Noh, J. H.; Yang, W. S.; Kim, Y. C.; Ryu, S.; Seo, J.; Seok, S. I.,

Compositional Engineering of Perovskite Materials for High-Performance Solar Cells. Nature 2015, 517, 476-480. 3.

Chirilă, A.; Reinhard, P.; Pianezzi, F.; Bloesch, P.; Uhl, A. R.; Fella, C.; Kranz, L.;

Keller, D.; Gretener, C.; Hagendorfer, H., Potassium-induced Surface Modification of Cu(In,Ga)Se2 Thin Films for High-Efficiency Solar Cells. Nat. Mater. 2013, 12, 1107-1111. 4.

De Wolf, S.; Holovsky, J.; Moon, S.-J.; Löper, P.; Niesen, B.; Ledinsky, M.; Haug, F.-J.;

Yum, J.-H.; Ballif, C., Organometallic Halide Perovskites: Sharp Optical Absorption Edge and Its Relation to Photovoltaic Performance. J. Phys. Chem. Lett. 2014, 5, 1035-1039.

ACS Paragon Plus Environment

15

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

5.

Page 16 of 20

Xing, G.; Mathews, N.; Sun, S.; Lim, S. S.; Lam, Y. M.; Gratzel, M.; Mhaisalkar, S.;

Sum, T. C., Long-Range Balanced Electron- and Hole-Transport Lengths in Organic-Inorganic CH3NH3PbI3. Science 2013, 342, 344-347. 6.

Stoumpos, C. C.; Malliakas, C. D.; Kanatzidis, M. G., Semiconducting Tin and Lead

Iodide Perovskites with Organic Cations: Phase Transitions, High Mobilities, and Near-Infrared Photoluminescent Properties. Inorg. Chem. 2013, 52, 9019-9038. 7.

Stranks, S. D.; Eperon, G. E.; Grancini, G.; Menelaou, C.; Alcocer, M. J.; Leijtens, T.;

Herz, L. M.; Petrozza, A.; Snaith, H. J., Electron-Hole Diffusion Lengths Exceeding 1 Micrometer in an Organometal Trihalide Perovskite Absorber. Science 2013, 342, 341-344. 8.

Zhao,

Y.;

Nardes,

A.

M.;

Zhu,

K.,

Solid-State Mesostructured

Perovskite

CH3NH3PbI3Solar Cells: Charge Transport, Recombination, and Diffusion Length. J. Phys. Chem. Lett. 2014, 5, 490-494. 9.

Kulkarni, S. A.; Baikie, T.; Boix, P. P.; Yantara, N.; Mathews, N.; Mhaisalkar, S., Band-

Gap Tuning of Lead Halide Perovskites Using a Sequential Deposition Process. J. Mater. Chem. A 2014, 2, 9221-9225. 10. Hao, F.; Stoumpos, C. C.; Chang, R. P.; Kanatzidis, M. G., Anomalous Band Gap Behavior in Mixed Sn and Pb Perovskites Enables Broadening of Absorption Spectrum in Solar Cells. J. Am. Chem. Soc. 2014, 136, 8094-8099. 11. Eperon, G. E.; Stranks, S. D.; Menelaou, C.; Johnston, M. B.; Herz, L. M.; Snaith, H. J., Formamidinium Lead Trihalide: a Broadly Tunable Perovskite for Efficient Planar Heterojunction Solar Cells. Energy Environ. Sci. 2014, 7, 982-988.

ACS Paragon Plus Environment

16

Page 17 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

12. Lee, M. M.; Teuscher, J.; Miyasaka, T.; Murakami, T. N.; Snaith, H. J., Efficient Hybrid Solar Cells Based on Meso-Superstructured Organometal Halide Perovskites. Science 2012, 338, 643-647. 13. Ball, J. M.; Lee, M. M.; Hey, A.; Snaith, H. J., Low-Temperature Processed MesoSuperstructured to Thin-Film Perovskite Solar Cells. Energy Environ. Sci. 2013, 6, 1739-1743. 14. Zhao, Y.; Zhu, K., CH3NH3Cl-Assisted One-Step Solution Growth of CH3NH3PbI3: Structure, Charge-Carrier Dynamics, and Photovoltaic Properties of Perovskite Solar Cells. J. Phys. Chem. C 2014, 118 , 9412-9418. 15. Colella, S.; Mosconi, E.; Fedeli, P.; Listorti, A.; Gazza, F.; Orlandi, F.; Ferro, P.; Besagni, T.; Rizzo, A.; Calestani, G.; Gigli, G.; De Angelis, F.; Mosca, R., MAPbI3-xClx Mixed Halide Perovskite for Hybrid Solar Cells: The Role of Chloride as Dopant on the Transport and Structural Properties. Chem. Mater. 2013, 25, 4613-4618. 16. Yu, H.; Wang, F.; Xie, F.; Li, W.; Chen, J.; Zhao, N., The Role of Chlorine in the Formation Process of “CH3NH3PbI3-xClx” Perovskite. Adv. Funct. Mater. 2014, 24, 7102-7108. 17. Mosconi, E.; Ronca, E.; De Angelis, F., First-Principles Investigation of the TiO2/Organohalide Perovskites Interface: The Role of Interfacial Chlorine. J. Phys. Chem. Lett. 2014, 5, 2619-2625. 18. Shi, J.; Luo, Y.; Wei, H.; Luo, J.; Dong, J.; Lv, S.; Xiao, J.; Xu, Y.; Zhu, L.; Xu, X.; Wu, H.; Li, D.; Meng, Q., Modified Two-Step Deposition Method for High-Efficiency TiO2/CH3NH3PbI3 Heterojunction Solar Cells. ACS Appl. Mater. Interfaces 2014, 6, 9711-9718.

ACS Paragon Plus Environment

17

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 20

19. Yongzhen, W.; Islam, A.; Yang, X.; Qin, C.; Liu, J.; Zhang, K.; Peng, W.; Han, L., Retarding the Crystallization of PbI2 for Highly Reproducible Planar-Structured Perovskite Solar Cells via Sequential Deposition. Energy Environ. Sci. 2014, 7, 2934-2938. 20. Jiang, Q.; Sheng, X.; Li, Y.; Feng, X.; Xu, T., Rutile TiO2 Nanowire-Based Perovskite Solar Cells. Chem. Commun. 2014, 50, 14720-14723. 21. Burschka, J.; Pellet, N.; Moon, S.-J.; Humphry-Baker, R.; Gao, P.; Nazeeruddin, M. K.; Grätzel, M., Sequential Deposition as a Route to High-Performance Perovskite-Sensitized Solar Cells. Nature 2013, 499, 316-319. 22. Liu, D.; Yang, J.; Kelly, T. L., Compact Layer Free Perovskite Solar Cells with 13.5% Efficiency. J. Am. Chem. Soc. 2014, 136, 17116-17122. 23. Sun, S.; Salim, T.; Mathews, N.; Duchamp, M.; Boothroyd, C.; Xing, G.; Sum, T. C.; Lam, Y. M., The Origin of High Efficiency in Low-Temperature Solution-Processable Bilayer Organometal Halide Hybrid Solar Cells. Energy Environ. Sci. 2014, 7, 399-407. 24. Zhu, G.; Lin, T.; Lü, X.; Zhao, W.; Yang, C.; Wang, Z.; Yin, H.; Liu, Z.; Huang, F.; Lin, J., Black Brookite Titania with High Solar Absorption and Excellent Photocatalytic Performance. J. Mater. Chem. A 2013, 1, 9650-9653. 25. King, R. R.; Bhusari, D.; Boca, A.; Larrabee, D.; Liu, X. Q.; Hong, W.; Fetzer, C. M.; Law, D. C.; Karam, N. H., Band Gap-Voltage Offset and Energy Production in Next-Generation Multijunction Solar Cells. Prog. Photovolt: Res. Appl. 2011, 19, 797-812.

ACS Paragon Plus Environment

18

Page 19 of 20

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

26. Chen, Q.; Zhou, H.; Hong, Z.; Luo, S.; Duan, H. S.; Wang, H. H.; Liu, Y.; Li, G.; Yang, Y., Planar Heterojunction Perovskite Solar Cells via Vapor-Assisted Solution Process. J. Am. Chem. Soc. 2014, 136, 622-625. 27. Hao, F.; Stoumpos, C. C.; Liu, Z.; Chang, R. P.; Kanatzidis, M. G., Controllable Perovskite Crystallization at a Gas-Solid Interface for Hole Conductor-Free Solar Cells with Steady Power Conversion Efficiency over 10%. J. Am. Chem. Soc. 2014, 136, 16411-16419. 28. Joshi, U. A.; Maggard, P. A., CuNb3O8: A p-Type Semiconducting Metal Oxide Photoelectrode. J. Phys. Chem. Lett. 2012, 3, 1577-1581. 29. Boschloo, G. K.; Goossens, A., Electron Trapping in Porphyrin-Sensitized Porous Nanocrystalline TiO2 Electrodes. J. Phys. Chem. 1996, 100, 19489-19494.

ACS Paragon Plus Environment

19

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 20

Table of Contents

ACS Paragon Plus Environment

20